This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Boundary feedback stabilization of quasilinear hyperbolic systems with partially dissipative structure

Ke Wang  Zhiqiang Wang  Wancong Yao Department of Mathematics, Donghua University, Shanghai 201620, China. E-mail: [email protected]. School of Mathematical Sciences and Shanghai Key Laboratory for Contemporary Applied Mathematics, Fudan University, Shanghai 200433, China. E-mail: [email protected]. School of Mathematical Sciences, Fudan University, Shanghai 200433, China. E-mail: [email protected].
(March 26, 2020)
Abstract

In this paper, we study the boundary feedback stabilization of a quasilinear hyperbolic system with partially dissipative structure. Thanks to this structure, we construct a suitable Lyapunov function which leads to the exponential stability to the equilibrium of the H2H^{2} solution. As an application, we also obtain the feedback stabilization for the Saint-Venant-Exner model under physical boundary conditions.

2010 Mathematics Subject Classification.  35L50,  93D15,  93D30,  35Q35

Key Words. Quasilinear hyperbolic system, feedback stabilization, Lyapunov function, Saint-Venant-Exner model

1 . Introduction and main results

Many models in physics, mechanics and other fields, including gas and fluid dynamics for instance, are described as hyperbolic equations. Control problems, particularly the stability and stabilization problems, of hyperbolic systems have been widely studied for decades (see [2, 3] and the references therein).

One classical approach to establish the asymptotic stability of hyperbolic system is the characteristic method. In the framework of C1C^{1}-solution, dissipative boundary conditions that lead to exponential stability of quasilinear hyperbolic systems without source terms have been found in [8, 15].

Another important approach to design boundary feedback controls is the Backstepping method. It has been used to stabilize exponentially the inhomogeneous quasilinear hyperbolic system in H2H^{2} norm (see [6, 12]). One can refer to [14] for many successful examples about feedback stabilization with this approach.

The third powerful approach is the Lyapunov function method. A strict Lyapunov function is introduced in [16] to achieve the exponential stability of a class of symmetric linear hyperbolic systems. Similar Lyapunov functions are used for quasilinear homogeneous hyperbolic systems in the framework of H2H^{2}-solution in [4].

If the hyperbolic system is inhomogeneous, the Lyapunov function approach can still be applied (see for instance [7, 9]). However, the nonzero source term change a lot the stability properties. With a source term, a simple quadratic Lyapunov function ensuring exponential stability for the L2L^{2} norm (or HpH^{p} norm) does not always exist no matter what the boundary conditions are. In [1], the authors study a linear 2×22\times 2 hyperbolic system and found a necessary and sufficient condition for simple quadratic Lyapunov function. Later in Chapter 6 of [2], the authors give a sufficient (but a priori non-necessary) condition such that the exponential stability of the system for the HpH^{p} norm with p2p\geq 2 is achieved. We refer to [10] for a relevant result in C1C^{1}-norm or CpC^{p}-norm. Naturally, these conditions all include one interior condition which requires a good coupling structure of the hyperbolic system, compared to the homogeneous case. However, as mentioned in their papers, this interior condition (typically a differential matrix inequality) is not straightforward to be checked in a specific model.

Different from the above, Herty and Yong study the boundary feedback stabilization of one-dimensional linear hyperbolic systems with a relaxation term in [11]. The key assumption is a structural stability condition which is introduced from [17] and is satisfies in many physical models. Later, in [18], Yong shows that under this structural stability condition, the boundary feedback stabilization result is also available for a class of one-dimensional linear hyperbolic system with vanishing eigenvalues.

Motivated by [11], in this paper, we consider a one-dimensional quasilinear hyperbolic system with the same relaxation structure. Thanks to the partial dissipation in the structural stability condition, we establish the local exponential stability of this nonlinear system for the H2H^{2}-norm. The main strategy is to construct a strict Lyapunov function together with a perturbation argument based on linear approximation. Compared to the result in [11], we provide an explicit sufficient condition on the gains of stabilizing boundary feedback control. As an application, we also obtain the boundary feedback stabilization of the Saint-Venant-Exner model proposed in [13] under physical boundary conditions.

Precisely, we are concerned with the boundary feedback stabilization of the following one-dimensional quasilinear hyperbolic system

Ut+𝐀(U)Ux=𝐐(U),t(0,),x(0,1)\displaystyle U_{t}+\mathbf{A}(U)U_{x}=\mathbf{Q}(U),\quad t\in(0,\infty),x\in(0,1) (1.1)

where U=(u1,,un)TU=(u_{1},\cdots,u_{n})^{T} is the unknown vector function of (t,x)(t,x), 𝐀:nn,n()\mathbf{A}:\mathbb{R}^{n}\mapsto\mathcal{M}_{n,n}(\mathbb{R}) is a smooth matrix function and 𝐐:nn\mathbf{Q}:\mathbb{R}^{n}\mapsto\mathbb{R}^{n} is a smooth vector function.

Let UnU^{*}\in\mathbb{R}^{n} be an equilibrium of (1.1), i.e.,

𝐐(U)=0.\displaystyle\mathbf{Q}(U^{*})=0. (1.2)

Without loss of generality, we may assume U=0U^{*}=0, otherwise one can consider UUU-U^{*} as the unknown functions.

We first assume that the system (1.1) is hyperbolic in a neighborhood of U=0U=0, i.e., the matrix 𝐀(U)\mathbf{A}(U) has nn real eigenvalues

𝚲r(U)<0<𝚲s(U)(r=1,,m;s=m+1,,n),\displaystyle\mathbf{\Lambda}_{r}(U)<0<\mathbf{\Lambda}_{s}(U)\quad(r=1,\cdots,m;\ s=m+1,\cdots,n), (1.3)

and it has a complete set of left eigenvectors 𝐋i(U)=(𝐋i1(U),,𝐋in(U)),(i=1,,n)\mathbf{L}_{i}(U)=(\mathbf{L}_{i1}(U),\cdots,\mathbf{L}_{in}(U)),\ (i=1,\cdots,n), i.e.,

𝐋i(U)𝐀(U)=𝚲i(U)𝐋i(U)(i=1,,n).\displaystyle\mathbf{L}_{i}(U)\mathbf{A}(U)=\mathbf{\Lambda}_{i}(U)\mathbf{L}_{i}(U)\quad(i=1,\cdots,n). (1.4)

Let

𝐋(U)=(𝐋1(U)𝐋n(U)) and 𝚲(U)=(𝚲(U)00𝚲+(U))\displaystyle\mathbf{L}(U)=\left(\begin{array}[]{ll}\mathbf{L}_{1}(U)\\ \vdots\\ \mathbf{L}_{n}(U)\end{array}\right)\text{\quad and \quad}\mathbf{\Lambda}(U)=\left(\begin{array}[]{cc}\mathbf{\Lambda}_{-}(U)&0\\ 0&\mathbf{\Lambda}_{+}(U)\end{array}\right) (1.10)

where

𝚲(U)=diag{𝚲1(U),,𝚲m(U)}, and 𝚲+(U)=diag{𝚲m+1(U),,𝚲n(U)}.\displaystyle\mathbf{\Lambda}_{-}(U)=\text{diag}\{\mathbf{\Lambda}_{1}(U),\cdots,\mathbf{\Lambda}_{m}(U)\},\text{\quad and \quad}\mathbf{\Lambda}_{+}(U)=\text{diag}\{\mathbf{\Lambda}_{m+1}(U),\cdots,\mathbf{\Lambda}_{n}(U)\}. (1.11)

Then

𝐋(U)𝐀(U)=𝚲(U)𝐋(U).\displaystyle\mathbf{L}(U)\mathbf{A}(U)=\mathbf{\Lambda}(U)\mathbf{L}(U). (1.12)

It is easy to see that system (1.1) is hyperbolic if and only if there is a symmetric positive definite matrix 𝐀0(U)\mathbf{A}_{0}(U), such that

𝐀0(U)𝐀(U)=𝐀T(U)𝐀0(U).\displaystyle\mathbf{A}_{0}(U)\mathbf{A}(U)=\mathbf{A}^{T}(U)\mathbf{A}_{0}(U). (1.13)

Then it follows that

(𝐋1(0))T𝐀0(0)𝐋1(0)𝚲(0)=𝚲(0)(𝐋1(0))T𝐀0(0)𝐋1(0).\displaystyle(\mathbf{L}^{-1}(0))^{T}\mathbf{A}_{0}(0)\mathbf{L}^{-1}(0)\mathbf{\Lambda}(0)=\mathbf{\Lambda}(0)(\mathbf{L}^{-1}(0))^{T}\mathbf{A}_{0}(0)\mathbf{L}^{-1}(0). (1.14)

Consequently, there exist two symmetric positive definite matrices 𝐗1(0)m,m()\mathbf{X}_{1}(0)\in\mathcal{M}_{m,m}(\mathbb{R}) and 𝐗2(0)nm,nm()\mathbf{X}_{2}(0)\in\mathcal{M}_{n-m,n-m}(\mathbb{R}) such that

(𝐋1(0))T𝐀0(0)𝐋1(0)=(𝐗1(0)00𝐗2(0)).\displaystyle(\mathbf{L}^{-1}(0))^{T}\mathbf{A}_{0}(0)\mathbf{L}^{-1}(0)=\left(\begin{array}[]{cc}\mathbf{X}_{1}(0)&0\\ 0&\mathbf{X}_{2}(0)\end{array}\right). (1.17)

Moreover, we assume the system possesses the following partially dissipative structure in a neighborhood of U=0U=0:

There exist invertible matrices 𝐏(U)n,n()\mathbf{P}(U)\in\mathcal{M}_{n,n}(\mathbb{R}) and 𝐒(U)r,r()\mathbf{S}(U)\in\mathcal{M}_{r,r}(\mathbb{R}) with 0<rn0<r\leq n, such that

𝐏(U)𝐐U(U)=(000𝐒(U))𝐏(U),\displaystyle\mathbf{P}(U)\mathbf{Q}_{U}(U)=\left(\begin{array}[]{cc}0&0\\ 0&\mathbf{S}(U)\end{array}\right)\mathbf{P}(U), (1.20)
𝐀0(U)𝐐U(U)+𝐐UT(U)𝐀0(U)𝐏T(U)(000𝐈r)𝐏(U).\displaystyle\mathbf{A}_{0}(U)\mathbf{Q}_{U}(U)+\mathbf{Q}_{U}^{T}(U)\mathbf{A}_{0}(U)\leq-\mathbf{P}^{T}(U)\left(\begin{array}[]{cc}0&0\\ 0&\mathbf{I}_{r}\end{array}\right)\mathbf{P}(U). (1.23)

Here 𝐐U(U)\mathbf{Q}_{U}(U) stands for the Jacobian matrix of 𝐐\mathbf{Q} with respect to UU, IrI_{r} denotes the r×rr\times r identity matrix. Let us point out that the above assumptions (1.20) and (1.23) are called structural stability conditions in [17, 11], which are commonly satisfied in lots of physical models.

Let

ξ(t,x)=𝐋(0)U(t,x)\displaystyle\xi(t,x)=\mathbf{L}(0)U(t,x) (1.24)

be the linearized diagonal variable and denote

ξ(t,x)=(ξ(t,x)ξ+(t,x))\displaystyle\xi(t,x)=\left(\begin{array}[]{ll}\xi_{-}(t,x)\\ \xi_{+}(t,x)\end{array}\right) (1.27)

where ξ(t,x)m\xi_{-}(t,x)\in\mathbb{R}^{m} and ξ+(t,x)nm\xi_{+}(t,x)\in\mathbb{R}^{n-m}. According to the theory on the well-posedness of the quasilinear hyperbolic system, the typical boundary conditions are given as follows

(ξ+(t,0)ξ(t,1))=𝐊(ξ+(t,1)ξ(t,0)),t(0,),\displaystyle\left(\begin{array}[]{ll}\xi_{+}(t,0)\\ \xi_{-}(t,1)\end{array}\right)=\mathbf{K}\left(\begin{array}[]{ll}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right),\quad t\in(0,\infty), (1.32)

where the feedback matrix

𝐊=(𝐊00𝐊01𝐊10𝐊11),\displaystyle\mathbf{K}=\left(\begin{array}[]{ll}\mathbf{K}_{00}&\mathbf{K}_{01}\\ \mathbf{K}_{10}&\mathbf{K}_{11}\end{array}\right), (1.35)

𝐊00nm,nm()\mathbf{K}_{00}\in\mathcal{M}_{n-m,n-m}(\mathbb{R}), 𝐊01nm,m()\mathbf{K}_{01}\in\mathcal{M}_{n-m,m}(\mathbb{R}), 𝐊10m,nm()\mathbf{K}_{10}\in\mathcal{M}_{m,n-m}(\mathbb{R}) and 𝐊11m,m()\mathbf{K}_{11}\in\mathcal{M}_{m,m}(\mathbb{R}) are all matrices with constant elements.

Finally, the initial condition is prescribed as

U(0,x)=U0(x),x(0,1),\displaystyle U(0,x)=U_{0}(x),\quad x\in(0,1), (1.36)

with U0H2((0,1);n)U_{0}\in H^{2}((0,1);\mathbb{R}^{n}) in a neighborhood of U=0U=0.

Regarding the well-posedness of the solutions to the problem (1.1), (1.32) and (1.36), we have the following proposition

Proposition 1.1.

There exists δ0>0\delta_{0}>0 such that, for every U0H2((0,1);n)U_{0}\in H^{2}((0,1);\mathbb{R}^{n}) satisfying

U0H2((0,1);n)δ0,\displaystyle||U_{0}||_{H^{2}((0,1);\mathbb{R}^{n})}\leq\delta_{0}, (1.37)

and the C1C^{1} compatibility conditions at the points (t,x)=(0,0)(t,x)=(0,0), (0,1)(0,1), the problem (1.1), (1.32) and (1.36) has a unique maximal classical solution

UC0([0,T);H2((0,1);n))\displaystyle U\in C^{0}([0,T);H^{2}((0,1);\mathbb{R}^{n})) (1.38)

with T(0,+]T\in(0,+\infty]. Moreover, if

U(t,)H2((0,1);n)δ0,t[0,T),\displaystyle||U(t,\cdot)||_{H^{2}((0,1);\mathbb{R}^{n})}\leq\delta_{0},\quad\forall t\in[0,T), (1.39)

then T=+T=+\infty.

Our main result is the following theorem.

Theorem 1.1.

Assume that the hyperbolic system (1.1) has no vanishing characteristic speed and it possesses the partially dissipative structure, i.e., (1.3), (1.13), (1.20) and (1.23) hold. Let 𝐊\mathbf{K} be chosen such that matrices

(𝐗2(0)𝚲+(0)00𝐗1(0)𝚲(0))𝐊T(𝐗2(0)𝚲+(0)00𝐗1(0)𝚲(0))𝐊\displaystyle\left(\begin{array}[]{cc}\mathbf{X}_{2}(0)\mathbf{\Lambda}_{+}(0)&0\\ 0&-\mathbf{X}_{1}(0)\mathbf{\Lambda}_{-}(0)\end{array}\right)-\mathbf{K}^{T}\left(\begin{array}[]{cc}\mathbf{X}_{2}(0)\mathbf{\Lambda}_{+}(0)&0\\ 0&-\mathbf{X}_{1}(0)\mathbf{\Lambda}_{-}(0)\end{array}\right)\mathbf{K} (1.44)

and

(e𝚲+(0)𝚲+(0)00𝚲(0))𝐊T(𝚲+(0)00e𝚲(0)𝚲(0))𝐊\displaystyle\left(\begin{array}[]{cc}e^{-\mathbf{\Lambda}_{+}(0)}\mathbf{\Lambda}_{+}(0)&0\\ 0&-\mathbf{\Lambda}_{-}(0)\end{array}\right)-\mathbf{K}^{T}\left(\begin{array}[]{cc}\mathbf{\Lambda}_{+}(0)&0\\ 0&-e^{-\mathbf{\Lambda}_{-}(0)}\mathbf{\Lambda}_{-}(0)\end{array}\right)\mathbf{K} (1.49)

are both positive definite. Then, the closed-loop system (1.1), (1.32) and (1.36) is locally exponentially stable for the H2H^{2}-norm, i.e., there exist positive constants δ\delta, CC and ν\nu, such that the solution to the system (1.1), (1.32) and (1.36) satisfies

U(t,)H2((0,1);n)CeνtU0H2((0,1);n),t[0,+),\displaystyle||U(t,\cdot)||_{H^{2}((0,1);\mathbb{R}^{n})}\leq Ce^{-\nu t}||U_{0}||_{H^{2}((0,1);\mathbb{R}^{n})},\quad t\in[0,+\infty), (1.50)

provided that

U0H2((0,1);n)δ||U_{0}||_{H^{2}((0,1);\mathbb{R}^{n})}\leq\delta (1.51)

and the C1C^{1} compatibility conditions are satisfied at (t,x)=(0,0)(t,x)=(0,0) and (0,1)(0,1).

Remark 1.1.

Theorem 1.1 still holds if the assumption (1.20) is extended to a more general case that

𝐏(U)𝐐U(U)𝐏1(U)=(𝐒11(U)𝐒12(U)𝐒21(U)𝐒22(U)),\displaystyle\mathbf{P}(U)\mathbf{Q}_{U}(U)\mathbf{P}^{-1}(U)=\left(\begin{array}[]{cc}\mathbf{S}_{11}(U)&\mathbf{S}_{12}(U)\\ \mathbf{S}_{21}(U)&\mathbf{S}_{22}(U)\end{array}\right), (1.54)

where 𝐒22(U)r,r()\mathbf{S}_{22}(U)\in\mathcal{M}_{r,r}(\mathbb{R}) with 0<rn0<r\leq n is an invertible matrix, |𝐒11(0)||\mathbf{S}_{11}(0)|_{\infty} and |𝐒21(0)||\mathbf{S}_{21}(0)|_{\infty} are sufficiently small.

Remark 1.2.

Theorem 1.1 still holds if the assumption (1.23) is extended to a more general case that

𝐀0(U)𝐐U(U)+𝐐UT(U)𝐀0(U)𝐏T(U)(000𝐑(U))𝐏(U),\displaystyle\mathbf{A}_{0}(U)\mathbf{Q}_{U}(U)+\mathbf{Q}_{U}^{T}(U)\mathbf{A}_{0}(U)\leq-\mathbf{P}^{T}(U)\left(\begin{array}[]{cc}0&0\\ 0&\mathbf{R}(U)\end{array}\right)\mathbf{P}(U), (1.57)

where 𝐑(U)r,r()\mathbf{R}(U)\in\mathcal{M}_{r,r}(\mathbb{R}) is a symmetric positive definite matrix.

Remark 1.3.

The conditions on the feedback matrix (1.44) and (1.49) are satisfied provided that |𝐊||\mathbf{K}|_{\infty} is sufficiently small. Particularly, if 𝐊\mathbf{K} is chosen as (1.35) where

𝐊00=κ+𝐈nm,𝐊11=κ𝐈m,𝐊01=0,𝐊10=0,\displaystyle\mathbf{K}_{00}=\kappa_{+}\mathbf{I}_{n-m},\quad\mathbf{K}_{11}=\kappa_{-}\mathbf{I}_{m},\quad\mathbf{K}_{01}=0,\quad\mathbf{K}_{10}=0,

with two constants κ+\kappa_{+} and κ\kappa_{-} satisfying

κ+2<exp(maxs=m+1,,n𝚲s(0))andκ2<exp(minr=1,,m𝚲r(0)).\displaystyle\kappa_{+}^{2}<\exp\Big{(}-\max\limits_{s=m+1,\cdots,n}\mathbf{\Lambda}_{s}(0)\Big{)}\quad\text{and}\quad\kappa_{-}^{2}<\exp\Big{(}\min\limits_{r=1,\cdots,m}\mathbf{\Lambda}_{r}(0)\Big{)}.
Remark 1.4.

Let us emphasize that the partially dissipative structure (1.20) and (1.23) combined with the dissipative boundary conditions (1.44) and (1.49) can be included by the interior and boundary stability conditions proposed in [2, Theorem 6.6]. However, the interior conditions on stability are typically differential matrix inequality, while the conditions proposed in this paper are all algebraic conditions which are more straightforward to be checked.

Remark 1.5.

It is also worthy of mentioning that the stability conditions (both interior and boundary conditions) for the nonlinear hyperbolic systems depend on the topology and in particular that the stability in H2H^{2} norm does not imply the stability in C1C^{1} norm (see [5]).

The paper is organized as follows: in Section 2, we introduce a new quasilinear hyperbolic system with a simpler structure under a transformation of the unknown functions. Then in Section 3, we construct a weighted H2H^{2}-Lyapunov function to prove the exponential stability of the new system which implies immediately Theorem 1.1. The proofs of related lemmas will be given in Section 4. Finally, in Section 5, the main result is applied to the Saint-Venant-Exner model for moving water on an open canal under physical boundary conditions.

2 Transformation of the system

In this section, we introduce a new hyperbolic system with a partially dissipative but simpler structure under a transformation of the unknown functions. In this way, the exponential stability of the original system is reduced to that of the new system.

Let

V=𝐏(0)U.\displaystyle V=\mathbf{P}(0)U. (2.1)

Then, the system (1.1) can be reduced to

Vt+A(V)Vx=B(V),\displaystyle V_{t}+A(V)V_{x}=B(V), (2.2)

where

A(V)=𝐏(0)𝐀(𝐏1(0)V)𝐏1(0) and B(V)=𝐏(0)𝐐(𝐏1(0)V).\displaystyle A(V)=\mathbf{P}(0)\mathbf{A}(\mathbf{P}^{-1}(0)V)\mathbf{P}^{-1}(0)\text{\quad and \quad}B(V)=\mathbf{P}(0)\mathbf{Q}(\mathbf{P}^{-1}(0)V). (2.3)

Clearly, V=0V=0 is an equilibrium of (2.2) and the Jacobian matrix of BB with respect to VV at the equilibrium can be calculated as

BV(0)=𝐏(0)𝐐U(0)𝐏1(0).B_{V}(0)=\mathbf{P}(0)\mathbf{Q}_{U}(0)\mathbf{P}^{-1}(0). (2.4)

Let

L(V)=𝐋(𝐏1(0)V)𝐏1(0) and Λ(V)=𝚲(𝐏1(0)V).\displaystyle L(V)=\mathbf{L}(\mathbf{P}^{-1}(0)V)\mathbf{P}^{-1}(0)\text{\quad and \quad}\Lambda(V)=\mathbf{\Lambda}(\mathbf{P}^{-1}(0)V). (2.5)

It is easy to check that L(V)L(V) is the matrix composed of the left eigenvectors of A(V)A(V), i.e.,

L(V)A(V)=Λ(V)L(V),\displaystyle L(V)A(V)=\Lambda(V)L(V), (2.6)

which implies that system (2.2) is a hyperbolic system without vanishing characteristic speeds. Obviously, we have

Λ(0)=𝚲(0)=diag{𝚲1(0),,𝚲n(0)}.\displaystyle\Lambda(0)=\mathbf{\Lambda}(0)=\text{diag}\{\mathbf{\Lambda}_{1}(0),\cdots,\mathbf{\Lambda}_{n}(0)\}. (2.7)

Let

A0(V)=(𝐏1(0))T𝐀0(𝐏1(0)V)𝐏1(0).\displaystyle A_{0}(V)=(\mathbf{P}^{-1}(0))^{T}\mathbf{A}_{0}(\mathbf{P}^{-1}(0)V)\mathbf{P}^{-1}(0). (2.8)

Obviously, A0(V)A_{0}(V) is a symmetric positive definite matrix satisfying

A0(V)A(V)=AT(V)A0(V).\displaystyle A_{0}(V)A(V)=A^{T}(V)A_{0}(V). (2.9)

Thanks to (2.4) and (2.8), the partially dissipative structure (1.20)-(1.23) for the original system (1.1) implies the following partially dissipative but simpler structure for system (2.2) at the equilibrium V=0V=0.

BV(0)=(000𝐒(0)),\displaystyle B_{V}(0)=\left(\begin{array}[]{cc}0&0\\ 0&\mathbf{S}(0)\end{array}\right), (2.12)
A0(0)BV(0)+BVT(0)A0(0)(000𝐈r).\displaystyle A_{0}(0)B_{V}(0)+B^{T}_{V}(0)A_{0}(0)\leq-\left(\begin{array}[]{cc}0&0\\ 0&\mathbf{I}_{r}\end{array}\right). (2.15)

According to the structure (2.12) and (2.15), we write V(t,x)V(t,x) as

V(t,x)=(v1(t,x)v2(t,x))with v1nr,v2r\displaystyle V(t,x)=\left(\begin{array}[]{c}v_{1}(t,x)\\ v_{2}(t,x)\end{array}\right)\quad\text{with \ }v_{1}\in\mathbb{R}^{n-r},\ v_{2}\in\mathbb{R}^{r} (2.18)

for further use.

From (1.24) and (2.5), we can see that the linear diagonal variables ξ(t,x)\xi(t,x) now becomes

ξ(t,x)=L(0)V(t,x),\displaystyle\xi(t,x)=L(0)V(t,x), (2.19)

with (1.27), which implies that the boundary conditions are still given by (1.32).

The initial condition for the variable VV is given by

V(0,x)=V0(x)𝐏(0)U0(x),x(0,1).\displaystyle V(0,x)=V_{0}(x)\triangleq\mathbf{P}(0)U_{0}(x),\quad x\in(0,1). (2.20)

In order to prove Thereom 1.1, it suffices to establish the H2H^{2}-stabilization for the system (2.2), (2.20) and (1.32).

3 Proof of Theorem 1.1

In this section, we can find suitable conditions on the feedback matrix 𝐊\mathbf{K} such that the closed-loop system (2.2), (2.20) and (1.32) is exponentially stable in H2(0,1)H^{2}(0,1)-norm. Then Theorem 1.1 follows immediately.

Let V0V_{0} with small H2((0,1);n)H^{2}((0,1);\mathbb{R}^{n}) norm be such that the C1C^{1} compatibility conditions at (t,x)=(0,0)(t,x)=(0,0) and (0,1)(0,1) are satisfied. Let also VC0([0,T),H2((0,1);n))V\in C^{0}([0,T),H^{2}((0,1);\mathbb{R}^{n})) be the maximal classical solution of the problem (2.2), (2.20) and (1.32). We remark that we only prove the stabilization result for smooth solutions while the conclusion follows easily from an density and continuity arguments for distributed solutions.

Motivated by [2] and [11], we construct a weighted Lyapunov function as follows:

𝕃(t)𝕃0(t)+𝕃1(t)+𝕃2(t)\displaystyle\mathbb{L}(t)\triangleq\mathbb{L}_{0}(t)+\mathbb{L}_{1}(t)+\mathbb{L}_{2}(t) (3.1)

with

𝕃0(t)01VT(αA0(V)+LT(V)e𝚲(0)xL(V))Vdx,\displaystyle\mathbb{L}_{0}(t)\triangleq\int^{1}_{0}V^{T}\Big{(}\alpha A_{0}(V)+L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)\Big{)}V\,\mathrm{d}x, (3.2)
𝕃1(t)01VtT(αA0(V)+LT(V)e𝚲(0)xL(V))Vtdx,\displaystyle\mathbb{L}_{1}(t)\triangleq\int^{1}_{0}V_{t}^{T}\Big{(}\alpha A_{0}(V)+L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)\Big{)}V_{t}\,\mathrm{d}x, (3.3)
𝕃2(t)01VttT(αA0(V)+LT(V)e𝚲(𝟎)xL(V))Vttdx,\displaystyle\mathbb{L}_{2}(t)\triangleq\int^{1}_{0}V_{tt}^{T}\Big{(}\alpha A_{0}(V)+L^{T}(V)e^{-\mathbf{\Lambda(0)}x}L(V)\Big{)}V_{tt}\,\mathrm{d}x, (3.4)

where e𝚲(0)x=diag{e𝚲1(0)x,,e𝚲n(0)x}e^{-\mathbf{\Lambda}(0)x}=\text{diag}\{e^{-\mathbf{\Lambda}_{1}(0)x},\cdots,e^{-\mathbf{\Lambda}_{n}(0)x}\}, α>0\alpha>0 is a constant to be chosen later.

For the simplicity of statements, we denote the L2(0,1)\|\cdot\|_{L^{2}(0,1)} norm as \|\cdot\|, C0([0,1])\|\cdot\|_{C^{0}([0,1])} norm as ||0|\cdot|_{0}, C1([0,1])\|\cdot\|_{C^{1}([0,1])} norm as ||1|\cdot|_{1}.

By definition of the Lyapunov function 𝕃(t)\mathbb{L}(t), 𝕃(t)\mathbb{L}(t) is equivalent to the energy V2+Vt2+Vtt2||V||^{2}+||V_{t}||^{2}+||V_{tt}||^{2} if |V(t,)|0|V(t,\cdot)|_{0} is small. On the other hand, Differentiation of system (2.2) with respect to tt and xx gives that

Vtt+A(V)Vtx=BV(V)Vt(A(V)Vt)Vx,\displaystyle V_{tt}+A(V)V_{tx}=B_{V}(V)V_{t}-(A^{\prime}(V)V_{t})V_{x}, (3.5)
Vtx+A(V)Vxx=BV(V)Vx(A(V)Vx)Vx,\displaystyle V_{tx}+A(V)V_{xx}=B_{V}(V)V_{x}-(A^{\prime}(V)V_{x})V_{x}, (3.6)

in which A(V)Vt,A(V)VxA^{\prime}(V)V_{t},A^{\prime}(V)V_{x} are matrices with entries aij(V)VVt,aij(V)VVx\frac{\partial a_{ij}(V)}{\partial V}V_{t},\frac{\partial a_{ij}(V)}{\partial V}V_{x} respectively. Then it is easy to see that if 𝕃(t)\mathbb{L}(t) is equivalent to the energy V(t,)H2((0,1);n)2||V(t,\cdot)||^{2}_{H^{2}((0,1);\mathbb{R}^{n})} if |V(t,)|1|V(t,\cdot)|_{1} is small.

Next, we turn to estimate the time derivative of 𝕃(t)\mathbb{L}(t). For this purpose, we can establish the following lemmas with the assumptions (1.44)-(1.49) on the feedback matrix 𝐊\mathbf{K}. The proof of the lemmas will be given in Section 4.

Lemma 3.1.

There exist positive constants α0\alpha_{0}, β0\beta_{0}, γ0\gamma_{0} and δ0\delta_{0} independent of VV such that if |V(t,)|0δ0|V(t,\cdot)|_{0}\leq\delta_{0},

𝕃0(t)α0v12+(β0α)v22+γ001(|V|3+|V|2|Vt|)dx,\displaystyle\mathbb{L}_{0}^{\prime}(t)\leq-\alpha_{0}||v_{1}||^{2}+(\beta_{0}-\alpha)||v_{2}||^{2}+\gamma_{0}\int^{1}_{0}(|V|^{3}+|V|^{2}|V_{t}|)\,\mathrm{d}x, (3.7)

where v1,v2v_{1},v_{2} is defined in (2.18).

Lemma 3.2.

There exist positive constants α1\alpha_{1}, β1\beta_{1}, γ1\gamma_{1} and δ1\delta_{1} independent of VV such that if |V(t,)|0δ1|V(t,\cdot)|_{0}\leq\delta_{1},

𝕃1(t)α1v1t2+(β1α)v2t2+γ101(|Vt|3+|V||Vt|2)dx.\displaystyle\mathbb{L}_{1}^{\prime}(t)\leq-\alpha_{1}||v_{1t}||^{2}+(\beta_{1}-\alpha)||v_{2t}||^{2}+\gamma_{1}\int^{1}_{0}(|V_{t}|^{3}+|V||V_{t}|^{2})\,\mathrm{d}x. (3.8)
Lemma 3.3.

There exist positive constants α2\alpha_{2}, β2\beta_{2}, γ2\gamma_{2} and δ2\delta_{2} independent of VV such that if |V(t,)|1δ2|V(t,\cdot)|_{1}\leq\delta_{2},

𝕃2(t)α2v1tt2+(β2α)v2tt2+γ201(|Vt||Vtt|2+|V||Vtt|2+|Vt|2|Vtt|)dx.\displaystyle\mathbb{L}_{2}^{\prime}(t)\leq-\alpha_{2}||v_{1tt}||^{2}+(\beta_{2}-\alpha)||v_{2tt}||^{2}+\gamma_{2}\int^{1}_{0}(|V_{t}||V_{tt}|^{2}+|V||V_{tt}|^{2}+|V_{t}|^{2}|V_{tt}|)\,\mathrm{d}x. (3.9)

With the help of Lemmas 3.1, 3.2 and 3.3, we are ready to prove Theorem 1.1.

Let the constants α>max{β0,β1,β2}\alpha>\max\{\beta_{0},\beta_{1},\beta_{2}\} and δ4min{δ0,δ1,δ2,δ3}\delta_{4}\leq\min\{\delta_{0},\delta_{1},\delta_{2},\delta_{3}\}. The combination of (3.7)-(3.9) yields that there exist positive constants β\beta and γ\gamma such that

𝕃(t)\displaystyle\mathbb{L}^{\prime}(t) β𝕃(t)+γ|V(t,)|1𝕃(t),\displaystyle\leq-\beta\mathbb{L}(t)+\gamma|V(t,\cdot)|_{1}\mathbb{L}(t), (3.10)

if |V(t,)|1δ4|V(t,\cdot)|_{1}\leq\delta_{4}.

Let δ5min{δ4,β2γ}\delta_{5}\triangleq\min\{\delta_{4},\frac{\beta}{2\gamma}\}. If we assume in a priori that |V(t,)|1δ5|V(t,\cdot)|_{1}\leq\delta_{5} for t(0,T)t\in(0,T), we get

𝕃(t)β2𝕃(t),t(0,T)\displaystyle\mathbb{L}^{\prime}(t)\leq-\frac{\beta}{2}\mathbb{L}(t),\quad t\in(0,T) (3.11)

which implies that 𝕃(t)\mathbb{L}(t) decays exponentially

𝕃(t)eβt2𝕃(0),t(0,T).\displaystyle\mathbb{L}(t)\leq e^{-\frac{\beta t}{2}}\mathbb{L}(0),\quad t\in(0,T). (3.12)

Using the equivalence of the energy V(t,)H2((0,1);n)2||V(t,\cdot)||^{2}_{H^{2}((0,1);\mathbb{R}^{n})} and 𝕃(t)\mathbb{L}(t), we obtain

V(t,)H2((0,1);n)C1eβt2V0H2((0,1);n),t[0,T)\displaystyle||V(t,\cdot)||_{H^{2}((0,1);\mathbb{R}^{n})}\leq C_{1}e^{-\frac{\beta t}{2}}||V_{0}||_{H^{2}((0,1);\mathbb{R}^{n})},\quad\forall t\in[0,T) (3.13)

for some constant C1>0C_{1}>0.

Note also the Sobolev inequality implies

|V(t,)|1C2V(t,)H2((0,1);n)C2C1V0H2((0,1);n),t[0,T).\displaystyle|V(t,\cdot)|_{1}\leq C_{2}||V(t,\cdot)||_{H^{2}((0,1);\mathbb{R}^{n})}\leq C_{2}C_{1}||V_{0}||_{H^{2}((0,1);\mathbb{R}^{n})},\quad\forall t\in[0,T). (3.14)

Let now

δ=δ5C2C1.\delta=\frac{\delta_{5}}{C_{2}C_{1}}. (3.15)

then the a priori estimate on |V(t,)|1δ5|V(t,\cdot)|_{1}\leq\delta_{5} indeed holds in [0,T)[0,T) if V0H2((0,1);n)δ||V_{0}||_{H^{2}((0,1);\mathbb{R}^{n})}\leq\delta. Therefore (3.13) follows immediately. According to Proposition 1.1, we finally conclude that the inequality (3.13) indeed holds for T=+T=+\infty.

Consequently, it follows from (2.1) that the solution UU to the problem (1.1), (1.32) and (1.36) is locally exponentially stable for the H2H^{2}-norm. This concludes the proof of Theorem 1.1.

4 Proof of the Lemmas

4.1 Proof of Lemma 2.1

We calculate the time-derivative of 𝕃0(t)\mathbb{L}_{0}(t) defined by (3.2),

𝕃0(t)=012(αVTA0(V)Vt+VTLT(V)e𝚲(0)xL(V)Vt)dx+𝒪(01|V|2|Vt|dx;|V(t,)|0).\displaystyle\mathbb{L}_{0}^{\prime}(t)=\int^{1}_{0}2\Big{(}\alpha V^{T}A_{0}(V)V_{t}+V^{T}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)V_{t}\Big{)}\,\mathrm{d}x+\mathcal{O}\Big{(}\int^{1}_{0}|V|^{2}|V_{t}|\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}. (4.1)

Here and hereafter 𝒪(X;Y)\mathcal{O}(X;Y) denotes the terms that for X0X\geq 0, Y0Y\geq 0, there exist C>0C>0 and ε>0\varepsilon>0 independent of VV, VtV_{t} and VttV_{tt}, satisfying

Yε|𝒪(X;Y)|CX.\displaystyle Y\leq\varepsilon\Rightarrow|\mathcal{O}(X;Y)|\leq CX. (4.2)

Substituting the system (2.2) into (4.1), we have

𝕃0(t)=𝕀0+𝕁0+𝒪(01|V|2|Vt|dx;|V(t,)|0)\displaystyle\mathbb{L}_{0}^{\prime}(t)=\mathbb{I}_{0}+\mathbb{J}_{0}+\mathcal{O}\Big{(}\int^{1}_{0}|V|^{2}|V_{t}|\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)} (4.3)

where

𝕀0012α(VTA0(V)B(V)VTA0(V)A(V)Vx)dx,\displaystyle\mathbb{I}_{0}\triangleq\int^{1}_{0}2\alpha\Big{(}V^{T}A_{0}(V)B(V)-V^{T}A_{0}(V)A(V)V_{x}\Big{)}\,\mathrm{d}x, (4.4)
𝕁0012(VTLT(V)e𝚲(0)xL(V)B(V)VTLT(V)e𝚲(0)xL(V)A(V)Vx)dx.\displaystyle\mathbb{J}_{0}\triangleq\int^{1}_{0}2\Big{(}V^{T}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)B(V)-V^{T}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)A(V)V_{x}\Big{)}\,\mathrm{d}x. (4.5)

Let’s first estimate the term 𝕀0\mathbb{I}_{0}. Using (2.9), (2.15) and integrations by parts, we have

𝕀0\displaystyle\mathbb{I}_{0} =01αVT(A0(0)BV(0)+BVT(0)A0(0))V(αVTA0(V)A(V)V)xdx\displaystyle=\int^{1}_{0}\alpha V^{T}\Big{(}A_{0}(0)B_{V}(0)+B^{T}_{V}(0)A_{0}(0)\Big{)}V-\Big{(}\alpha V^{T}A_{0}(V)A(V)V\Big{)}_{x}\,\mathrm{d}x
+𝒪(01(|V|3+|V|2|Vt|)dx;|V(t,)|0)\displaystyle\quad+\mathcal{O}\Big{(}\int^{1}_{0}(|V|^{3}+|V|^{2}|V_{t}|)\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}
αv22+[αVTA0(V)A(V)V]|01+𝒪(01(|V|3+|V|2|Vt|)dx;|V(t,)|0).\displaystyle\leq-\alpha||v_{2}||^{2}+\Big{[}-\alpha V^{T}A_{0}(V)A(V)V\Big{]}\Big{|}_{0}^{1}+\mathcal{O}\Big{(}\int^{1}_{0}(|V|^{3}+|V|^{2}|V_{t}|)\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}. (4.6)

Then we turn to estimate the term 𝕁0\mathbb{J}_{0}. Linear approximation together with (2.6) implies

𝕁0\displaystyle\mathbb{J}_{0} =012VTLT(0)e𝚲(0)xL(0)BV(0)Vdx01VTLT(0)e𝚲(0)x𝚲2(0)L(0)Vdx\displaystyle=\int^{1}_{0}2V^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}L(0)B_{V}(0)V\,\mathrm{d}x-\int^{1}_{0}V^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}^{2}(0)L(0)V\,\mathrm{d}x
01(VTLT(V)e𝚲(0)xΛ(V)L(V)V)xdx+𝒪(01(|V|3+|V|2|Vt|)dx;|V(t,)|0).\displaystyle\quad-\int^{1}_{0}\Big{(}V^{T}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}\Lambda(V)L(V)V\Big{)}_{x}\,\mathrm{d}x+\mathcal{O}\Big{(}\int^{1}_{0}(|V|^{3}+|V|^{2}|V_{t}|)\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}. (4.7)

Note that the matrix LT(0)e𝚲(0)xL(0)L^{T}(0)e^{-\mathbf{\Lambda}(0)x}L(0) is symmetric and positive definite. We denote it as following block matrix 𝐌\mathbf{M}

𝐌(x)(𝐌11(x)𝐌12(x)𝐌12T(x)𝐌22(x))\displaystyle\mathbf{M}(x)\triangleq\left(\begin{array}[]{cc}\mathbf{M}_{11}(x)&\mathbf{M}_{12}(x)\\ \mathbf{M}^{T}_{12}(x)&\mathbf{M}_{22}(x)\end{array}\right) (4.10)

according to the block matrix BV(0)B_{V}(0). Then by Cauchy-Schwarz inequality, we get for all ε>0\varepsilon>0, that

012VTLT(0)e𝚲(0)xL(0)BV(0)Vdx=\displaystyle\int^{1}_{0}2V^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}L(0)B_{V}(0)V\,\mathrm{d}x= 012v1T𝐌12(x)𝐒(0)v2+2v2T𝐌22(x)𝐒(0)v2dx\displaystyle\int^{1}_{0}2v_{1}^{T}\mathbf{M}_{12}(x)\mathbf{S}(0)v_{2}+2v_{2}^{T}\mathbf{M}_{22}(x)\mathbf{S}(0)v_{2}\,\mathrm{d}x
\displaystyle\leq εv12+Cεv22,\displaystyle\varepsilon||v_{1}||^{2}+C_{\varepsilon}||v_{2}||^{2}, (4.11)

where Cε>0C_{\varepsilon}>0 is constant depending on ε\varepsilon. Because LT(0)e𝚲(0)x𝚲2(0)L(0)L^{T}(0)e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}^{2}(0)L(0) is positive definite, there exists a constant c0>0c_{0}>0 such that

01VTLT(0)e𝚲(0)x𝚲2(0)L(0)Vdxc0V2.\displaystyle-\int^{1}_{0}V^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}^{2}(0)L(0)V\,\mathrm{d}x\leq-c_{0}||V||^{2}. (4.12)

Thus, it follows from (4.7),(4.1)-(4.12) and integrations by parts that

𝕁0\displaystyle\mathbb{J}_{0}\leq (εc0)v12+(Cεc0)v22+[VTLT(V)e𝚲(0)xΛ(V)L(V)V]|01\displaystyle(\varepsilon-c_{0})||v_{1}||^{2}+(C_{\varepsilon}-c_{0})||v_{2}||^{2}+\Big{[}-V^{T}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}\Lambda(V)L(V)V\Big{]}\Big{|}_{0}^{1}
+𝒪(01(|V|3+|V|2|Vt|)dx;|V(t,)|0).\displaystyle+\mathcal{O}\Big{(}\int^{1}_{0}(|V|^{3}+|V|^{2}|V_{t}|)\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}. (4.13)

Let ε=c02\varepsilon=\frac{c_{0}}{2}. Then combining (4.3), (4.6) and (4.1) yields that there exist positive constants α0\alpha_{0}, β0\beta_{0} independent of VV, such that

𝕃0(t)α0v12+(β0α)v22+𝔹0+𝒪(01(|V|3+|V|2|Vt|)dx;|V(t,)|0),\displaystyle\mathbb{L}_{0}^{\prime}(t)\leq-\alpha_{0}||v_{1}||^{2}+(\beta_{0}-\alpha)||v_{2}||^{2}+\mathbb{B}_{0}+\mathcal{O}\Big{(}\int^{1}_{0}(|V|^{3}+|V|^{2}|V_{t}|)\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}, (4.14)

where the boundary term

𝔹0\displaystyle\mathbb{B}_{0} =VT(t,x)(αA0(V)A(V)+LT(V)e𝚲(0)xΛ(V)L(V))V(t,x)|01.\displaystyle=-V^{T}(t,x)\Big{(}\alpha A_{0}(V)A(V)+L^{T}(V)e^{-\mathbf{\Lambda}(0)x}\Lambda(V)L(V)\Big{)}V(t,x)\Big{|}^{1}_{0}. (4.15)

It remains to estimate 𝔹0\mathbb{B}_{0}. Using (2.6), (2.19) and the linear approximation, we have

𝔹0\displaystyle\mathbb{B}_{0} =ξT(t,x)(α(L1(0))TA0(0)L1(0)𝚲(0)+e𝚲(0)x𝚲(0))ξ(t,x)|01\displaystyle=-\xi^{T}(t,x)\Big{(}\alpha(L^{-1}(0))^{T}A_{0}(0)L^{-1}(0)\mathbf{\Lambda}(0)+e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}(0)\Big{)}\xi(t,x)\Big{|}^{1}_{0}
+𝒪(|V(t,0)|3+|V(t,1)|3;|V(t,0)|+|V(t,1)|).\displaystyle\quad+\mathcal{O}\Big{(}|V(t,0)|^{3}+|V(t,1)|^{3};|V(t,0)|+|V(t,1)|\Big{)}. (4.16)

Noting (2.5) and (2.8), (1.17), we easily obtain that

(L1(0))TA0(0)L1(0)=(𝐋1(0))T𝐀0(0)𝐋1(0)=(𝐗1(0)00𝐗2(0)).\displaystyle(L^{-1}(0))^{T}A_{0}(0)L^{-1}(0)=(\mathbf{L}^{-1}(0))^{T}\mathbf{A}_{0}(0)\mathbf{L}^{-1}(0)=\left(\begin{array}[]{cc}\mathbf{X}_{1}(0)&0\\ 0&\mathbf{X}_{2}(0)\end{array}\right). (4.19)

Substituting the boundary condition (1.32) and (4.19) into (4.1), we thus get

𝔹0=(ξ+(t,1)ξ(t,0))T𝐆(ξ+(t,1)ξ(t,0))+𝒪(|V(t,0)|3+|V(t,1)|3;|V(t,0)|+|V(t,1)|)\displaystyle\mathbb{B}_{0}=-\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right)^{T}\mathbf{G}\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right)+\mathcal{O}\Big{(}|V(t,0)|^{3}+|V(t,1)|^{3};|V(t,0)|+|V(t,1)|\Big{)} (4.24)

where 𝐆\mathbf{G} is a symmetric matrix defined as

𝐆\displaystyle\mathbf{G}\triangleq α[(𝐗2(0)𝚲+(0)00𝐗1(0)𝚲(0))𝐊T(𝐗2(0)𝚲+(0)00𝐗1(0)𝚲(0))𝐊]\displaystyle\,\alpha\left[\left(\begin{array}[]{cc}\mathbf{X}_{2}(0)\mathbf{\Lambda}_{+}(0)&0\\ 0&-\mathbf{X}_{1}(0)\mathbf{\Lambda}_{-}(0)\end{array}\right)-\mathbf{K}^{T}\left(\begin{array}[]{cc}\mathbf{X}_{2}(0)\mathbf{\Lambda}_{+}(0)&0\\ 0&-\mathbf{X}_{1}(0)\mathbf{\Lambda}_{-}(0)\end{array}\right)\mathbf{K}\right] (4.29)
+[(e𝚲+(0)𝚲+(0)00𝚲(0))𝐊T(𝚲+(0)00e𝚲(0)𝚲(0))𝐊].\displaystyle+\left[\left(\begin{array}[]{cc}e^{-\mathbf{\Lambda}_{+}(0)}\mathbf{\Lambda}_{+}(0)&0\\ 0&-\mathbf{\Lambda}_{-}(0)\end{array}\right)-\mathbf{K}^{T}\left(\begin{array}[]{cc}\mathbf{\Lambda}_{+}(0)&0\\ 0&-e^{-\mathbf{\Lambda}_{-}(0)}\mathbf{\Lambda}_{-}(0)\end{array}\right)\mathbf{K}\right]. (4.34)

Using (2.19) and the boundary condition (1.32), we have

V(t,0)=L1(0)ξ(t,0)=L1(0)(0𝐈m𝐊00𝐊01)(ξ+(t,1)ξ(t,0))\displaystyle V(t,0)=L^{-1}(0)\xi(t,0)=L^{-1}(0)\left(\begin{array}[]{cc}0&\mathbf{I}_{m}\\ \mathbf{K}_{00}&\mathbf{K}_{01}\end{array}\right)\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right) (4.39)
V(t,1)=L1(0)ξ(t,1)=L1(0)(𝐊10𝐊11𝐈nm0)(ξ+(t,1)ξ(t,0))\displaystyle V(t,1)=L^{-1}(0)\xi(t,1)=L^{-1}(0)\left(\begin{array}[]{cc}\mathbf{K}_{10}&\mathbf{K}_{11}\\ \mathbf{I}_{n-m}&0\end{array}\right)\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right) (4.44)

Consequently, (4.24) becomes

𝔹0=(ξ+(t,1)ξ(t,0))T𝐆(ξ+(t,1)ξ(t,0))+𝒪((|V(t,0)|+|V(t,1)|)|(ξ+(t,1)ξ(t,0))|2;|V(t,0)|+|V(t,1)|).\displaystyle\mathbb{B}_{0}=-\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right)^{T}\mathbf{G}\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right)+\mathcal{O}\Big{(}(|V(t,0)|+|V(t,1)|)\left|\left(\begin{array}[]{c}\xi_{+}(t,1)\\ \xi_{-}(t,0)\end{array}\right)\right|^{2};|V(t,0)|+|V(t,1)|\Big{)}. (4.51)

Note that the assumptions (1.44)-(1.49) on the boundary feedback matrix 𝐊\mathbf{K} implies the symmetric matrix 𝐆\mathbf{G} is positive definite. Therefore, there exist δ0>0\delta_{0}>0 and γ0>0\gamma_{0}>0 such that the boundary term 𝔹00,\mathbb{B}_{0}\leq 0, and furthermore the estimate (3.7) holds if |V(t,)|0<δ0|V(t,\cdot)|_{0}<\delta_{0}. This concludes the proof of Lemma 3.1.

4.2 Proof of Lemma 2.2

By (3.3), the time-derivative of 𝕃1(t)\mathbb{L}_{1}(t) can be expressed as

𝕃1(t)\displaystyle\mathbb{L}_{1}^{\prime}(t) =012(αVtTA0(V)Vtt+VtTLT(V)e𝚲(0)xL(V)Vtt)dx+𝒪(01|Vt|3dx;|V(t,)|0).\displaystyle=\int^{1}_{0}2\Big{(}\alpha V_{t}^{T}A_{0}(V)V_{tt}+V^{T}_{t}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)V_{tt}\Big{)}\,\mathrm{d}x+\mathcal{O}\Big{(}\int^{1}_{0}|V_{t}|^{3}\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}. (4.52)

Substituting the term of VttV_{tt} derived from (3.5) into (4.52), we have

𝕃1(t)\displaystyle\mathbb{L}_{1}^{\prime}(t) =012αVtTA0(V)(BV(V)Vt(A(V)Vt)VxA(V)Vtx)dx\displaystyle=\int^{1}_{0}2\alpha V_{t}^{T}A_{0}(V)\Big{(}B_{V}(V)V_{t}-(A^{\prime}(V)V_{t})V_{x}-A(V)V_{tx}\Big{)}\,\mathrm{d}x
+012VtTLT(V)e𝚲(0)xL(V)(BV(V)Vt(A(V)Vt)VxA(V)Vtx)dx\displaystyle\quad+\int^{1}_{0}2V^{T}_{t}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)\Big{(}B_{V}(V)V_{t}-(A^{\prime}(V)V_{t})V_{x}-A(V)V_{tx}\Big{)}\,\mathrm{d}x
+𝒪(01|Vt|3dx;|V(t,)|0)\displaystyle\quad+\mathcal{O}\Big{(}\int^{1}_{0}|V_{t}|^{3}\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}

Thus, by using integrations by parts and some straightforward calculations, we get

𝕃1(t)\displaystyle\mathbb{L}_{1}^{\prime}(t) =01αVtT(A0(0)BV(0)+BV(0)TA0(0))VtVtTLT(0)e𝚲(0)x𝚲2(0)L(0)Vtdx\displaystyle=\int^{1}_{0}\alpha V_{t}^{T}\Big{(}A_{0}(0)B_{V}(0)+B_{V}(0)^{T}A_{0}(0)\Big{)}V_{t}-V_{t}^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}^{2}(0)L(0)V_{t}\,\mathrm{d}x
+012VtTLT(0)e𝚲(0)xL(0)BV(0)VtdxVtT(αA0(V)A(V)+LT(V)e𝚲(0)xΛ(V)L(V))Vt|01\displaystyle\quad+\int^{1}_{0}2V_{t}^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}L(0)B_{V}(0)V_{t}\,\mathrm{d}x-V_{t}^{T}\Big{(}\alpha A_{0}(V)A(V)+L^{T}(V)e^{-\mathbf{\Lambda}(0)x}\Lambda(V)L(V)\Big{)}V_{t}\Big{|}_{0}^{1}
+𝒪(01(|Vt|3+|V||Vt|2)dx;|V(t,)|0).\displaystyle\quad+\mathcal{O}\Big{(}\int^{1}_{0}(|V_{t}|^{3}+|V||V_{t}|^{2})\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}.

Similarly as the analysis of 𝕃0(t)\mathbb{L}_{0}^{\prime}(t) in the proof of Lemma 3.1, we obtain that there exist positive constants α1\alpha_{1} and β1\beta_{1}, such that

𝕃1(t)α1v1t2+(β1α)v2t2+𝔹1+𝒪(01(|Vt|3+|V||Vt|2)dx;|V(t,)|0),\displaystyle\mathbb{L}_{1}^{\prime}(t)\leq-\alpha_{1}||v_{1t}||^{2}+(\beta_{1}-\alpha)||v_{2t}||^{2}+\mathbb{B}_{1}+\mathcal{O}\Big{(}\int^{1}_{0}(|V_{t}|^{3}+|V||V_{t}|^{2})\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}, (4.53)

where the boundary term 𝔹1\mathbb{B}_{1} is

𝔹1\displaystyle\mathbb{B}_{1} =ξtT(t,x)(α(L1(0))TA0(0)L1(0)𝚲(0)+e𝚲(0)x𝚲(0))ξt(t,x)|01\displaystyle=-\xi^{T}_{t}(t,x)\Big{(}\alpha(L^{-1}(0))^{T}A_{0}(0)L^{-1}(0)\mathbf{\Lambda}(0)+e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}(0)\Big{)}\xi_{t}(t,x)\Big{|}^{1}_{0}
+𝒪((|V(t,0)||Vt(t,0)|2+|V(t,1)||Vt(t,1)|2);|V(t,0)|+|V(t,1)|).\displaystyle\quad+\mathcal{O}\Big{(}(|V(t,0)||V_{t}(t,0)|^{2}+|V(t,1)||V_{t}(t,1)|^{2});|V(t,0)|+|V(t,1)|\Big{)}. (4.54)

Taking the time-derivative (1.32) and (4.39)-(4.44), we can easily express (Vt(t,0),Vt(t,1))(V_{t}(t,0),V_{t}(t,1)) in terms of (ξ+t(t,1),ξt(t,0))(\xi_{+t}(t,1),\xi_{-t}(t,0)), thus the boundary term 𝔹1\mathbb{B}_{1} can be rewritten as

𝔹1\displaystyle\mathbb{B}_{1} =(ξ+t(t,1)ξt(t,0))T𝐆(ξ+t(t,1)ξt(t,0))\displaystyle=-\left(\begin{array}[]{c}\xi_{+t}(t,1)\\ \xi_{-t}(t,0)\end{array}\right)^{T}\mathbf{G}\left(\begin{array}[]{c}\xi_{+t}(t,1)\\ \xi_{-t}(t,0)\end{array}\right) (4.59)
+𝒪((|V(t,0)|+|V(t,1)|)|(ξ+t(t,1)ξt(t,0))|2;|V(t,0)|+|V(t,1)|).\displaystyle\quad+\mathcal{O}\Big{(}(|V(t,0)|+|V(t,1)|)\left|\left(\begin{array}[]{c}\xi_{+t}(t,1)\\ \xi_{-t}(t,0)\end{array}\right)\right|^{2};|V(t,0)|+|V(t,1)|\Big{)}. (4.62)

Note again with the assumptions (1.44)-(1.49) can imply that 𝐆\mathbf{G} is positive definite. Therefore, there exists δ1>0\delta_{1}>0 and γ1>0\gamma_{1}>0 such that 𝔹10,\mathbb{B}_{1}\leq 0, and furthermore the estimate (3.8) holds if |V(t,)|0<δ1|V(t,\cdot)|_{0}<\delta_{1}. The finishes the proof of Lemma 3.2.

4.3 Proof of Lemma 2.3

Calculating the time-derivative of 𝕃2(t)\mathbb{L}_{2}(t) gives

𝕃2(t)=012(αVttTA0(V)Vttt+VttTLT(V)e𝚲(0)xL(V)Vttt)dx+𝒪(01|Vt||Vtt|2dx;|V(t,)|0).\displaystyle\mathbb{L}_{2}^{\prime}(t)=\int^{1}_{0}2\Big{(}\alpha V_{tt}^{T}A_{0}(V)V_{ttt}+V^{T}_{tt}L^{T}(V)e^{-\mathbf{\Lambda}(0)x}L(V)V_{ttt}\Big{)}\,\mathrm{d}x+\mathcal{O}\Big{(}\int^{1}_{0}|V_{t}||V_{tt}|^{2}\,\mathrm{d}x;|V(t,\cdot)|_{0}\Big{)}. (4.63)

Differentiating system (3.5) with respect to tt and combining (2.2) and (3.5), we have,

Vttt+A(V)Vttx=BV(V)Vtt+(BV(V))tVt2(A(V)Vt)Vtx(A(V)Vt)tVx.\displaystyle V_{ttt}+A(V)V_{ttx}=B_{V}(V)V_{tt}+(B_{V}(V))_{t}V_{t}-2(A^{\prime}(V)V_{t})V_{tx}-(A^{\prime}(V)V_{t})_{t}V_{x}. (4.64)

Substituting the term of VtttV_{ttt} derived from (4.64) into 𝕃2(t)\mathbb{L}_{2}^{\prime}(t), we do integration by parts and linear approximation, as in the proof of Lemma 3.1 and 3.2, to deduce that

𝕃2(t)\displaystyle\mathbb{L}_{2}^{\prime}(t) =01αVttT(A0(0)BV(0)+BV(0)TA0(0))VttVttTLT(0)e𝚲(0)x𝚲2(0)L(0)Vttdx\displaystyle=\int^{1}_{0}\alpha V_{tt}^{T}\Big{(}A_{0}(0)B_{V}(0)+B_{V}(0)^{T}A_{0}(0)\Big{)}V_{tt}-V_{tt}^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}^{2}(0)L(0)V_{tt}\,\mathrm{d}x
+012VttTLT(0)e𝚲(0)xL(0)BV(0)VttdxVttT(αA0(V)A(V)+LT(V)e𝚲(0)xΛ(V)L(V))Vtt|01\displaystyle\quad+\int^{1}_{0}2V_{tt}^{T}L^{T}(0)e^{-\mathbf{\Lambda}(0)x}L(0)B_{V}(0)V_{tt}\,\mathrm{d}x-V_{tt}^{T}\Big{(}\alpha A_{0}(V)A(V)+L^{T}(V)e^{-\mathbf{\Lambda}(0)x}\Lambda(V)L(V)\Big{)}V_{tt}\Big{|}_{0}^{1}
+𝒪(01(|V||Vtt|2+|Vt||Vtt|2+|Vt|2|Vtt|dx;|V(t,)|1).\displaystyle\quad+\mathcal{O}\Big{(}\int^{1}_{0}(|V||V_{tt}|^{2}+|V_{t}||V_{tt}|^{2}+|V_{t}|^{2}|V_{tt}|\,\mathrm{d}x;|V(t,\cdot)|_{1}\Big{)}. (4.65)

Thanks to the partially dissipative structure (2.12)-(2.15) and the fact that LT(0)e𝚲(0)x𝚲2(0)L(0)L^{T}(0)e^{-\mathbf{\Lambda}(0)x}\mathbf{\Lambda}^{2}(0)L(0) is positive definite, there exist positive constants α2\alpha_{2} and β2\beta_{2} independent of VV such that

𝕃2(t)\displaystyle\mathbb{L}_{2}^{\prime}(t) α2v1tt2+(β2α)v2tt2+𝔹2\displaystyle\leq-\alpha_{2}||v_{1tt}||^{2}+(\beta_{2}-\alpha)||v_{2tt}||^{2}+\mathbb{B}_{2}
+𝒪(01(|V||Vtt|2+|Vt||Vtt|2+|Vt|2|Vtt|dx;|V(t,)|1)\displaystyle\quad+\mathcal{O}\Big{(}\int^{1}_{0}(|V||V_{tt}|^{2}+|V_{t}||V_{tt}|^{2}+|V_{t}|^{2}|V_{tt}|\,\mathrm{d}x;|V(t,\cdot)|_{1}\Big{)} (4.66)

where 𝔹2\mathbb{B}_{2} denotes the boundary term derived from integration by parts. Taking the second time-derivative of (1.32) and (4.39)-(4.44), we can rewrite the boundary term 𝔹2\mathbb{B}_{2} as

𝔹2\displaystyle\mathbb{B}_{2} =(ξ+tt(t,1)ξtt(t,0))T𝐆(ξ+tt(t,1)ξtt(t,0))\displaystyle=-\left(\begin{array}[]{c}\xi_{+tt}(t,1)\\ \xi_{-tt}(t,0)\end{array}\right)^{T}\mathbf{G}\left(\begin{array}[]{c}\xi_{+tt}(t,1)\\ \xi_{-tt}(t,0)\end{array}\right) (4.71)
+𝒪((|V(t,0)|+|V(t,1)|)|(ξ+tt(t,1)ξtt(t,0))|2;|V(t,0)|+|V(t,1)|).\displaystyle\quad+\mathcal{O}\Big{(}(|V(t,0)|+|V(t,1)|)\left|\left(\begin{array}[]{c}\xi_{+tt}(t,1)\\ \xi_{-tt}(t,0)\end{array}\right)\right|^{2};|V(t,0)|+|V(t,1)|\Big{)}. (4.74)

Note again that 𝐆\mathbf{G} is positive definite if the assumptions (1.44)-(1.49) hold. Therefore, there exists δ2>0\delta_{2}>0 and γ2>0\gamma_{2}>0 such that 𝔹20\mathbb{B}_{2}\leq 0 and furthermore the estimate (3.9) holds if |V(t,)|1<δ2|V(t,\cdot)|_{1}<\delta_{2}. This ends the proof of Lemma 3.3.

5 Application to Saint-Venant-Exner equations

We now consider the Saint-Venant-Exner equations for a moving bathymetry on a sloping channel with a rectangular cross-section:

Ht+VHx+HVx=0,Bt+aV2Vx=0,Vt+VVx+gHx+gBx=gSbCfV2H.\displaystyle\left.\begin{array}[]{lll}&H_{t}+VH_{x}+HV_{x}=0,\\ &B_{t}+aV^{2}V_{x}=0,\\ &V_{t}+VV_{x}+gH_{x}+gB_{x}=gS_{b}-C_{f}\frac{V^{2}}{H}.\end{array}\right. (5.4)

Here H=H(t,x)H=H(t,x) is the water depth, B=B(t,x)B=B(t,x) is the elevation of the sediment bed and V=V(t,x)>0V=V(t,x)>0 is the average velocity of water. Moreover, gg is the gravity constant, constant SbS_{b} is the bottom slope of the channel, constant CfC_{f} means the friction coefficient and constant aa is a parameter that includes porosity and viscosity effects on the sediment dynamics (see [13]).

Let (H,B,V)T(H_{*},B_{*},V_{*})^{T} with H>0,B>0H_{*}>0,B_{*}>0 and V>0V_{*}>0 be a constant equilibrium of (5.4), i.e.,

gSbH=Cf(V)2>0.\displaystyle gS_{b}H_{*}=C_{f}(V_{*})^{2}>0. (5.5)

Let U(h,b,v)TU\triangleq(h,b,v)^{T} be the deviations of the states:

h(t,x)=H(t,x)H,b(t,x)=B(t,x)B,v(t,x)=V(t,x)V.\displaystyle h(t,x)=H(t,x)-H_{*},\quad b(t,x)=B(t,x)-B_{*},\quad v(t,x)=V(t,x)-V_{*}. (5.6)

Then the Saint-Venant-Exner equations (5.4) can be rewritten in the form of (1.1) with

𝐀(U)=(v+V0h+H00a(v+V)2ggv+V),𝐐(U)=(00gSbCf(v+V)2h+H).\displaystyle\mathbf{A}(U)=\left(\begin{array}[]{ccc}v+V_{*}&0&h+H_{*}\\ 0&0&a(v+V_{*})^{2}\\ g&g&v+V_{*}\end{array}\right),\quad\mathbf{Q}(U)=\left(\begin{array}[]{c}0\\ 0\\ gS_{b}-C_{f}\frac{(v+V_{*})^{2}}{h+H_{*}}\end{array}\right). (5.13)

In order to apply Theorem 1.1, we will verify that the hyperbolic system satisfies the partially dissipative structure. For simplicity, we only show that (1.3), (1.13), (1.20) and (1.23) are satisfied at the equilibrium U=0U=0.

First, the matrix 𝐀(0)\mathbf{A}(0) has three eigenvalues λi\lambda_{i} (i=1,2,3)(i=1,2,3) satisfying

λ32Vλ2+(V2gaV2gH)λ+gaV3=0.\displaystyle\lambda^{3}-2V_{*}\lambda^{2}+(V_{*}^{2}-gaV_{*}^{2}-gH_{*})\lambda+gaV_{*}^{3}=0. (5.14)

Thus we get the following relations

λ1λ2λ3=gaV3,λ1+λ2+λ3=2V,λ1λ2+λ1λ3+λ2λ3=V2gaV2gH.\displaystyle\left.\begin{array}[]{lll}\lambda_{1}\lambda_{2}\lambda_{3}=-gaV_{*}^{3},\\ \lambda_{1}+\lambda_{2}+\lambda_{3}=2V_{*},\\ \lambda_{1}\lambda_{2}+\lambda_{1}\lambda_{3}+\lambda_{2}\lambda_{3}=V_{*}^{2}-gaV_{*}^{2}-gH_{*}.\end{array}\right. (5.18)

Here λ1\lambda_{1}, λ3\lambda_{3} are the characteristic velocities of the water flow and λ2\lambda_{2} is the characteristic velocity of the sediment motion. Therefore, (5.5) and (5.18) yield that matrix 𝐀(0)\mathbf{A}(0) has no vanishing eigenvalues with

λ1<0<λ2λ3.\displaystyle\lambda_{1}<0<\lambda_{2}\leq\lambda_{3}. (5.19)

Due to the fact that the sediment motion is much slower than the water flow, we can make the following reasonable assumptions that λ2\lambda_{2} is so small that

λ2<λ1and0<λ2<32V,\displaystyle\lambda_{2}<-\lambda_{1}\quad\text{and}\quad 0<\lambda_{2}<\frac{3}{2}V_{*}, (5.20)

which leads to the following relations

λ132V<0,λ232V<0,λ332V=12Vλ1λ2>0.\displaystyle\lambda_{1}-\frac{3}{2}V_{*}<0,\quad\lambda_{2}-\frac{3}{2}V_{*}<0,\quad\lambda_{3}-\frac{3}{2}V_{*}=\frac{1}{2}V_{*}-\lambda_{1}-\lambda_{2}>0. (5.21)

Next, we choose an invertible matrix 𝐏(0)\mathbf{P}(0) and a symmetric positive definite matrix 𝐀0(0)\mathbf{A}_{0}(0), such that (1.13), (1.20) and (1.23) are satisfied. Inspired by [11], we take

𝐏(0)=(100010V2H01) and 𝐀0(0)=(4gH+2agV24H2g2HV2Hg2H3g2aV20V2H01).\displaystyle\mathbf{P}(0)=\left(\begin{array}[]{ccc}1&0&0\\ 0&1&0\\ -\frac{V_{*}}{2H_{*}}&0&1\end{array}\right)\text{\quad and \quad}\mathbf{A}_{0}(0)=\left(\begin{array}[]{ccc}\frac{4gH_{*}+2agV_{*}^{2}}{4H_{*}^{2}}&-\frac{g}{2H_{*}}&-\frac{V_{*}}{2H_{*}}\\ -\frac{g}{2H_{*}}&\frac{3g}{2aV_{*}^{2}}&0\\ -\frac{V_{*}}{2H_{*}}&0&1\end{array}\right). (5.28)

According to (5.18) and (5.21), it is easy to verify that

det(𝐀0(0))=3g22aHV2+g22H23g8a2H2=gaH2V3i=13(λi32V)>0.\displaystyle\det(\mathbf{A}_{0}(0))=\frac{3g^{2}}{2aH_{*}V_{*}^{2}}+\frac{g^{2}}{2H_{*}^{2}}-\frac{3g}{8a^{2}H_{*}^{2}}=\frac{g}{aH_{*}^{2}V_{*}^{3}}\prod_{i=1}^{3}\Big{(}\lambda_{i}-\frac{3}{2}V_{*}\Big{)}>0. (5.29)

Therefore, 𝐀0(0)\mathbf{A}_{0}(0) is symmetric positive definite. It follows also that 𝐀0(0)𝐀(0)=𝐀T(0)𝐀0(0)\mathbf{A}_{0}(0)\mathbf{A}(0)=\mathbf{A}^{T}(0)\mathbf{A}_{0}(0).

Note that

𝐐U(0)=(000000CfV2H202CfVH).\displaystyle\mathbf{Q}_{U}(0)=\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ C_{f}\frac{{V_{*}}^{2}}{{H_{*}}^{2}}&0&-2C_{f}\frac{V_{*}}{H_{*}}\end{array}\right). (5.33)

Direct computations give that

𝐏(0)𝐐U(0)𝐏1(0)=(00000000CfVH)\displaystyle\mathbf{P}(0)\mathbf{Q}_{U}(0)\mathbf{P}^{-1}(0)=\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ 0&0&-C_{f}\frac{V_{*}}{H_{*}}\end{array}\right) (5.37)
(𝐏1(0))T(𝐀0(0)𝐐U(0)+𝐐UT(0)𝐀0(0))𝐏1(0)=(000000004CfVH).\displaystyle(\mathbf{P}^{-1}(0))^{T}\Big{(}\mathbf{A}_{0}(0)\mathbf{Q}_{U}(0)+\mathbf{Q}_{U}^{T}(0)\mathbf{A}_{0}(0)\Big{)}\mathbf{P}^{-1}(0)=\left(\begin{array}[]{ccc}0&0&0\\ 0&0&0\\ 0&0&-4C_{f}\frac{V_{*}}{H_{*}}\end{array}\right). (5.41)

i.e., the partially dissipative structure indeeds holds.

Let 𝐋i(0)\mathbf{L}_{i}(0) be the left eigenvectors corresponding to λi\lambda_{i} (i=1,2,3)(i=1,2,3), then we have

𝐋(0)=(𝐋1(0)𝐋2(0)𝐋3(0))=(gλ1Vgλ11gλ2Vgλ21gλ3Vgλ31)\displaystyle\mathbf{L}(0)=\left(\begin{array}[]{c}\mathbf{L}_{1}(0)\\ \mathbf{L}_{2}(0)\\ \mathbf{L}_{3}(0)\end{array}\right)=\left(\begin{array}[]{ccc}\frac{g}{\lambda_{1}-V_{*}}&\frac{g}{\lambda_{1}}&1\\ \frac{g}{\lambda_{2}-V_{*}}&\frac{g}{\lambda_{2}}&1\\ \frac{g}{\lambda_{3}-V_{*}}&\frac{g}{\lambda_{3}}&1\end{array}\right) (5.48)

and further

𝐋1(0)=(λ1H(λ1λ2)(λ1λ3)λ2H(λ2λ3)(λ2λ1)λ3H(λ3λ1)(λ3λ2)aV2(λ1V)(λ1λ2)(λ1λ3)aV2(λ2V)(λ2λ3)(λ2λ1)aV2(λ3V)(λ3λ1)(λ3λ2)λ1(λ1V)(λ1λ2)(λ1λ3)λ2(λ2V)(λ1λ2)(λ1λ3)λ3(λ3V)(λ3λ1)(λ3λ2)).\displaystyle\mathbf{L}^{-1}(0)=\left(\begin{array}[]{ccc}\frac{\lambda_{1}H_{*}}{(\lambda_{1}-\lambda_{2})(\lambda_{1}-\lambda_{3})}&\frac{\lambda_{2}H_{*}}{(\lambda_{2}-\lambda_{3})(\lambda_{2}-\lambda_{1})}&\frac{\lambda_{3}H_{*}}{(\lambda_{3}-\lambda_{1})(\lambda_{3}-\lambda_{2})}\vspace{2mm}\\ \frac{aV_{*}^{2}(\lambda_{1}-V_{*})}{(\lambda_{1}-\lambda_{2})(\lambda_{1}-\lambda_{3})}&\frac{aV_{*}^{2}(\lambda_{2}-V_{*})}{(\lambda_{2}-\lambda_{3})(\lambda_{2}-\lambda_{1})}&\frac{aV_{*}^{2}(\lambda_{3}-V_{*})}{(\lambda_{3}-\lambda_{1})(\lambda_{3}-\lambda_{2})}\vspace{2mm}\\ \frac{\lambda_{1}(\lambda_{1}-V_{*})}{(\lambda_{1}-\lambda_{2})(\lambda_{1}-\lambda_{3})}&\frac{\lambda_{2}(\lambda_{2}-V_{*})}{(\lambda_{1}-\lambda_{2})(\lambda_{1}-\lambda_{3})}&\frac{\lambda_{3}(\lambda_{3}-V_{*})}{(\lambda_{3}-\lambda_{1})(\lambda_{3}-\lambda_{2})}\end{array}\right). (5.52)

Consequently, we have

(𝐋1(0))T𝐀0(0)𝐋1(0)=(𝐗11000𝐗22000𝐗33)\displaystyle(\mathbf{L}^{-1}(0))^{T}\mathbf{A}_{0}(0)\mathbf{L}^{-1}(0)=\left(\begin{array}[]{ccc}\mathbf{X}_{11}&0&0\\ 0&\mathbf{X}_{22}&0\\ 0&0&\mathbf{X}_{33}\end{array}\right) (5.56)

where

𝐗11=λ1(λ132V)(λ1λ2)(λ1λ3),𝐗22=λ2(λ232V)(λ2λ1)(λ2λ3)𝐗33=λ3(λ332V)(λ3λ1)(λ3λ2).\displaystyle\mathbf{X}_{11}=\frac{\lambda_{1}(\lambda_{1}-\frac{3}{2}V^{*})}{(\lambda_{1}-\lambda_{2})(\lambda_{1}-\lambda_{3})},\quad\mathbf{X}_{22}=\frac{\lambda_{2}(\lambda_{2}-\frac{3}{2}V^{*})}{(\lambda_{2}-\lambda_{1})(\lambda_{2}-\lambda_{3})}\quad\mathbf{X}_{33}=\frac{\lambda_{3}(\lambda_{3}-\frac{3}{2}V^{*})}{(\lambda_{3}-\lambda_{1})(\lambda_{3}-\lambda_{2})}. (5.57)

For the Saint-Venant-Exner equations (5.4), we introduce the following physical boundary conditions

{b(t,0)=0,v(t,0)=k1h(t,0),v(t,1)=k2(h(t,1)+b(t,1)).\displaystyle\left\{\begin{array}[]{lll}b(t,0)=0,\\ v(t,0)=-k_{1}h(t,0),\\ v(t,1)=-k_{2}(h(t,1)+b(t,1)).\end{array}\right. (5.61)

where k1k_{1} and k2k_{2} are two feedback parameters.

Then the boundary conditions (5.61) can be rewritten in the form of its linearized diagonal variable ξ\xi defined by (1.24) as

(ξ2(t,0)ξ3(t,0)ξ1(t,1))=𝐊(ξ2(t,1)ξ3(t,1)ξ1(t,0)) with 𝐊=(00π2(k1)00π3(k1)χ2(k2)χ3(k2)0),\displaystyle\left(\begin{array}[]{ccc}\xi_{2}(t,0)\\ \xi_{3}(t,0)\\ \xi_{1}(t,1)\end{array}\right)=\mathbf{K}\left(\begin{array}[]{ccc}\xi_{2}(t,1)\\ \xi_{3}(t,1)\\ \xi_{1}(t,0)\end{array}\right)\text{\quad with \quad}\mathbf{K}=\left(\begin{array}[]{ccc}0&0&\pi_{2}(k_{1})\\ 0&0&\pi_{3}(k_{1})\\ \chi_{2}(k_{2})&\chi_{3}(k_{2})&0\end{array}\right), (5.71)

where πj\pi_{j} and χj\chi_{j} (j=2,3)(j=2,3) are the following quantities depending on k1k_{1} and k2k_{2}:

πj(k1)=λ1VλjVgk1(λjV)gk1(λ1V)(j=2,3),\displaystyle\pi_{j}(k_{1})=\frac{\lambda_{1}-V_{*}}{\lambda_{j}-V_{*}}\cdot\frac{g-k_{1}(\lambda_{j}-V_{*})}{g-k_{1}(\lambda_{1}-V_{*})}\quad(j=2,3), (5.72)
χ2(k2)=λ2(λ3λ1)(λ2V)λ1(λ3λ2)(λ1V)g+k2(λ2V)g+k2(λ1V),\displaystyle\chi_{2}(k_{2})=\frac{\lambda_{2}(\lambda_{3}-\lambda_{1})(\lambda_{2}-V_{*})}{\lambda_{1}(\lambda_{3}-\lambda_{2})(\lambda_{1}-V_{*})}\cdot\frac{g+k_{2}(\lambda_{2}-V_{*})}{g+k_{2}(\lambda_{1}-V_{*})}, (5.73)
χ3(k2)=λ3(λ1λ2)(λ3V)λ1(λ3λ2)(λ1V)g+k2(λ3V)g+k2(λ1V).\displaystyle\chi_{3}(k_{2})=\frac{\lambda_{3}(\lambda_{1}-\lambda_{2})(\lambda_{3}-V_{*})}{\lambda_{1}(\lambda_{3}-\lambda_{2})(\lambda_{1}-V_{*})}\cdot\frac{g+k_{2}(\lambda_{3}-V_{*})}{g+k_{2}(\lambda_{1}-V_{*})}. (5.74)

Consequently, the dissipative boundary conditions (1.44) and (1.49) yield

π22(k1)𝐗22𝐗11λ2|λ1|+π32(k1)𝐗33𝐗11λ3|λ1|1,\displaystyle\pi_{2}^{2}(k_{1})\frac{\mathbf{X}_{22}}{\mathbf{X}_{11}}\frac{\lambda_{2}}{|\lambda_{1}|}+\pi_{3}^{2}(k_{1})\frac{\mathbf{X}_{33}}{\mathbf{X}_{11}}\frac{\lambda_{3}}{|\lambda_{1}|}\leq 1, (5.75)
χ22(k2)𝐗11𝐗22|λ1|λ2+χ32(k2)𝐗11𝐗33|λ1|λ31\displaystyle\chi_{2}^{2}(k_{2})\frac{\mathbf{X}_{11}}{\mathbf{X}_{22}}\frac{|\lambda_{1}|}{\lambda_{2}}+\chi_{3}^{2}(k_{2})\frac{\mathbf{X}_{11}}{\mathbf{X}_{33}}\frac{|\lambda_{1}|}{\lambda_{3}}\leq 1 (5.76)

and

π22(k1)λ2|λ1|+π32(k1)λ3|λ1|1,\displaystyle\pi_{2}^{2}(k_{1})\frac{\lambda_{2}}{|\lambda_{1}|}+\pi_{3}^{2}(k_{1})\frac{\lambda_{3}}{|\lambda_{1}|}\leq 1, (5.77)
χ22(k2)eλ2λ1|λ1|λ2+χ32(k2)eλ3λ1|λ1|λ31.\displaystyle\chi_{2}^{2}(k_{2})e^{\lambda_{2}-\lambda_{1}}\frac{|\lambda_{1}|}{\lambda_{2}}+\chi_{3}^{2}(k_{2})e^{\lambda_{3}-\lambda_{1}}\frac{|\lambda_{1}|}{\lambda_{3}}\leq 1. (5.78)

Let

βj=max{𝐗jj𝐗11,1},ηj=max{𝐗11𝐗jj,eλjλ1},(j=2,3).\displaystyle\beta_{j}=\max\Big{\{}\frac{\mathbf{X}_{jj}}{\mathbf{X}_{11}},1\Big{\}},\ \eta_{j}=\max\Big{\{}\frac{\mathbf{X}_{11}}{\mathbf{X}_{jj}},e^{\lambda_{j}-\lambda_{1}}\Big{\}},\quad(j=2,3). (5.79)

Then, (5.75), (5.76), (5.77) and (5.78) are satisfied if

π22(k1)β2λ2|λ1|+π32(k1)β3λ3|λ1|1,\displaystyle\pi_{2}^{2}(k_{1})\beta_{2}\frac{\lambda_{2}}{|\lambda_{1}|}+\pi_{3}^{2}(k_{1})\beta_{3}\frac{\lambda_{3}}{|\lambda_{1}|}\leq 1, (5.80)
χ22(k2)η2|λ1|λ2+χ32(k2)η3|λ1|λ31.\displaystyle\chi_{2}^{2}(k_{2})\eta_{2}\frac{|\lambda_{1}|}{\lambda_{2}}+\chi_{3}^{2}(k_{2})\eta_{3}\frac{|\lambda_{1}|}{\lambda_{3}}\leq 1. (5.81)

Finally we conclude by Theorem 1.1 that

Theorem 5.1.

If the boundary feedback parameters k1k_{1} and k2k_{2} satisfy (5.80) and (5.81), respectively, then the constant equilibrium (H,B,V)T(H_{*},B_{*},V_{*})^{T} of the Saint-Venant-Exner system (5.4), (5.61) is locally exponentially stable for the H2H^{2}-norm.

Acknowledgements

The authors were partially supported by the National Science Foundation of China (No. 11971119) and the Fundamental Research Funds for the Central Universities (No. 2232020D-41).

References

  • [1] Georges Bastin and Jean-Michel Coron. On boundary feedback stabilization of non-uniform linear 2×22\times 2 hyperbolic systems over a bounded interval. Systems Control Lett., 60(11):900–906, 2011.
  • [2] Georges Bastin and Jean-Michel Coron. Stability and Boundary Stabilization of 1-D Hyperbolic Systems, volume 88 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser/Springer, [Cham], 2016. Subseries in Control.
  • [3] Jean-Michel Coron. Control and nonlinearity, volume 136 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2007.
  • [4] Jean-Michel Coron, Georges Bastin, and Brigitte d’Andréa Novel. Dissipative boundary conditions for one-dimensional nonlinear hyperbolic systems. SIAM J. Control Optim., 47(3):1460–1498, 2008.
  • [5] Jean-Michel Coron and Hoai-Minh Nguyen. Dissipative boundary conditions for nonlinear 1-D hyperbolic systems: sharp conditions through an approach via time-delay systems. SIAM J. Math. Anal., 47(3):2220–2240, 2015.
  • [6] Jean-Michel Coron, Rafael Vazquez, Miroslav Krstic, and Georges Bastin. Local exponential H2H^{2} stabilization of a 2×22\times 2 quasilinear hyperbolic system using backstepping. SIAM J. Control Optim., 51(3):2005–2035, 2013.
  • [7] Markus Dick, Martin Gugat, and Günter Leugering. A strict H1H^{1}-Lyapunov function and feedback stabilization for the isothermal Euler equations with friction. Numer. Algebra Control Optim., 1(2):225–244, 2011.
  • [8] J. M. Greenberg and Tatsien Li. The effect of boundary damping for the quasilinear wave equation. J. Differential Equations, 52(1):66–75, 1984.
  • [9] Martin Gugat, Günter Leugering, and Ke Wang. Neumann boundary feedback stabilization for a nonlinear wave equation: a strict H2H^{2}-Lyapunov function. Math. Control Relat. Fields, 7(3):419–448, 2017.
  • [10] Amaury Hayat. Boundary stability of 1-D nonlinear inhomogeneous hyperbolic systems for the C1C^{1} norm. SIAM J. Control Optim., 57(6):3603–3638, 2019.
  • [11] Michael Herty and Wen-An Yong. Feedback boundary control of linear hyperbolic systems with relaxation. Automatica J. IFAC, 69:12–17, 2016.
  • [12] Long Hu, Rafael Vazquez, Florent Di Meglio, and Miroslav Krstic. Boundary exponential stabilization of 1-dimensional inhomogeneous quasi-linear hyperbolic systems. SIAM J. Control Optim., 57(2):963–998, 2019.
  • [13] Justin Hudson and Peter K. Sweby. Formulations for numerically approximating hyperbolic systems governing sediment transport. J. Sci. Comput., 19(1-3):225–252, 2003. Special issue in honor of the sixtieth birthday of Stanley Osher.
  • [14] Miroslav Krstic and Andrey Smyshlyaev. Boundary control of PDEs, volume 16 of Advances in Design and Control. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2008. A course on backstepping designs.
  • [15] Tatsien Li. Global Classical Solutions for Quasilinear Hyperbolic Systems, volume 32 of RAM: Research in Applied Mathematics. Masson, Paris; John Wiley & Sons, Ltd., Chichester, 1994.
  • [16] Cheng-Zhong Xu and Gauthier Sallet. Exponential stability and transfer functions of processes governed by symmetric hyperbolic systems. ESAIM Control Optim. Calc. Var., 7:421–442, 2002.
  • [17] Wen-An Yong. Singular perturbations of first-order hyperbolic systems with stiff source terms. J. Differential Equations, 155(1):89–132, 1999.
  • [18] Wen-An Yong. Boundary stabilization of hyperbolic balance laws with characteristic boundaries. Automatica J. IFAC, 101:252–257, 2019.